Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Synthesis and luminescence properties of Ba3Lu(PO4)3:Sm3+ phosphor for white light-emitting diodes

Open Access Open Access

Abstract

A series of Ba3Lu(PO4)3:Sm3+ phosphors were prepared by traditional high temperature solid-state reaction methods. The site-preferred occupancy of Sm3+ in Ba3Lu(PO4)3 and the luminescence properties of Ba3Lu(PO4)3:Sm3+ were studied combined with X-ray diffraction, photoluminescence excitation (PLE) spectra, and emission (PL) spectra as well as temperature-dependent PL and decay curves. The PL intensity is improved with increasing Sm3+ content and the optimal dopant content is 0.05. The temperature-dependent PL spectra indicate that the emission intensity decreases with the temperature because of the enhancement of the non-radiative transition. The results indicate that these reddish-orange emitting phosphors could be for potential applications in w-LEDs.

© 2017 Optical Society of America

1. Introduction

In recent years, white light-emitting diodes (w-LEDs) have attracted increasing attentions from engineers and scientists in the solid-state luminescence field owing to the merits of being environmentally friendly, their long lifetime, energy-saving qualities, and high luminous efficiency [1–3]. At present, the most common method is to employ the yellow-emitting YAG:Ce3+ phosphor with blue InGaN chips to produce the practical white light emission [4]. However, commercial w-LED lamps are deficient in the red region, resulting in a lowly satisfactory color-rendering index and poor light performance. A suitable red-emitting phosphor for near ultraviolet (NUV) phosphor converted LEDs should have many potential applications, due to their excellent color rendering index, high color tolerance and high conversion efficiency into visible light [5]. Therefore, it is urgent to find new red phosphors that can be excited by NUV-LED chips for the fabrication of w-LEDs.

Phosphors based on a phosphate host is one of the most important luminescent materials and rare earth doped phosphates usually have excellent thermal stability, ideal charge stabilization and low sintering temperature, which have attracted more and more attention [6]. To our knowledge, eulytite-type orthophosphate with the general formula AII3MIII(PO4)3 (A = alkaline earth or Pb; M = trivalent rare earth, Bi, or V) have been reported [7–10]. The results show that eulytite-type orthophosphate are considered as suitable host matrices for phosphors due to the characteristics of high physical and chemical stability, water-resistant property, stable crystal structure, and excellent optical properties.

Owing to the great requirement for the red phosphors used in developing solid state lighting, such as LEDs, many different types of hosts have been developed, such as K2TiF6:Mn4+, MZnOS:Mn2+ (M = Ca, Ba), CuInS, BaNb2O6:Sm3+ and so on [11–14]. Among them, Sm3+ doped red phosphors played an important role in these fields. As is known, Sm3+ ion is an important activator for its narrow emission band and high quantum efficiency. The emission and excitation of Sm3+ belong to the 4f-4f transition, besides it has good red light emission excited by NUV light, thus causing extensive researches [15]. To the best of our knowledge, no paper has been published on the preparation of Sm3+ doped Ba3Lu(PO4)3 phosphor. So we synthesized Ba3Lu(PO4)3:Sm3+ phosphor via the high temperature solid-state reaction method. The high temperature solid-state reaction method is very important in the preparation of fluorescent materials. This method is not only used early but also widely used in real life. The samples prepared via high temperature solid-state reaction method has some outstanding features compared to other methods such as high brightness, low cost, simple preparation process and good quality powder. And the most important is that this method is easy to realize industrialized production [16].The luminescence properties of this phosphor were studied by photoluminescence excitation (PLE) spectrum, photoluminescence emission (PL) spectrum, concentration quenching, and decay curve analyses.

2. Experimental

The stoichiometric amounts of raw materials, BaCO3 (A.R.), Lu2O3 (A.R.), (NH4)2HPO4 (A.R.) and Sm2O3 (99.99%) were thoroughly mixed and ground in an agate mortar. After the stoichiometric amounts of the starting materials were mixed thoroughly, the mixtures were fired in a muffle furnace at 650 °C for 5 h, reground, and finally fired at 1250 °C for 3 h in air. After calcining, the samples were cooled to room temperature and ground thoroughly into powder for measurement.

The phase structure of the as-prepared phosphor was recorded by an X-ray Powder diffraction spectroscopy (XRD, Bruker D2, Karlsruhe, Germany) with Cu Kα radiation (λ = 1.5406 Å) operated at 30 kV tube voltage and 10 mA tube current. The morphology of the samples was determined by using a field-emission scanning electron microscope equipped with an energy-dispersive spectrometer (FE-SEM, S-4800, Hitachi, Japan). The amounts of Sm elements in precursor solution and filtrate were analyzed by inductively coupled plasma-atomic emission spectroscopy (ICP-AES) using a Perkin Elmer Optima 8000 instrument. The PL and PLE spectra of the samples were measured on a fluorescence spectrophotometer (F-7000, Hitachi, Japan) with a photomultiplier tube operating at 400 V, and 150 W Xe lamp used as an excitation source. Diffuse reflection spectra were recorded using a UV–Vis–NIR spectrophotometer (UV-3700, SHIMADZU) and white BaSO4 powder was used as a reference standard. The luminescence decay curve was obtained using a spectrofluorometer (HORIBA, JOBIN YVON FL3-21), and the 370 nm pulse laser radiation was used as the excitation source. The temperature-dependence luminescence properties were measured on the same spectrophotometer, which was combined with a self-made heating attachment and a computer-controlled electric furnace (Tianjin Orient KOJI Co. Ltd, TAP-02).

3. Results and discussion

3.1 Structural studies

The structure of Ba3Lu(PO4)3 compound is illustrated in Fig. 1(a). Ba3Lu(PO4)3 reportedly crystallizes to a cubic crystal structure with a space group of I-43d (no.220) and lattice parameter values of a = 10.4760 Å, and V = 1149.71 Å3 [17,18]. As shown in Fig. 1(b), in this structure, the Ba2+/Lu3+ pairs of cations are disordered on a single crystallographic site, whereas the oxygen atoms disordered over two orientations (O1 and O2 with 35% and 65% occupancy factors, respectively) of the phosphate groups are distributed over three partially occupied sites [19]. As is known, the ionic radii of Lu3+ (r = 0.085 nm with CN = 9) and Sm3+ (r = 0.098 nm with CN = 9) are very close [20]. Given the identical electric charges of Lu3+ and Sm3+ and their similar radii, doped Sm3+ ions can substitute for Lu3+ sites.

 figure: Fig. 1

Fig. 1 (a) Structural view of Ba3Lu(PO4)3; (b) The coordination environment of Ba2+/Lu3+ site.

Download Full Size | PDF

A SEM is a powerful magnification tool that utilizes focused beams of electrons to obtain information. The high-resolution, 3D images produced by SEM provide topographical, morphological and compositional information, which makes them invaluable in a variety of science and industry applications [21]. Figure 2(a) and Fig. 2(b) shows the surface morphology of Ba3Lu(PO4)3:0.05Sm3+ phosphor at different magnifications. From the SEM images it is observed that the morphology of the powders is basically flaky polycrystalline constituted by microcrystalline agglomerated particles. In the current study, the SEM images of the phosphor reveal the average particle sizes of 2–10 μm. This is a suitable size for fabrication of solid-state lighting devices.

 figure: Fig. 2

Fig. 2 SEM images of Ba3Lu(PO4)3:0.05Sm3+ sample under different magnification (a): 2.2 k × ; (b) 3.5 k × .

Download Full Size | PDF

A series of XRD patterns of Ba3Lu(PO4)3:Sm3+ phosphors with different doping contents as well as the Joint Committee for Powder Diffraction Standard (JCPDS) file 43-0212 [Ba3Lu(PO4)3] are illustrated in Fig. 3. It can be seen that no detectable impurity phase is observed in the obtained samples when the doping contents is below 0.09, indicating that a single-phase phosphor is obtained and the doped Sm3+ ions do not change the host crystal lattice. In the meantime, the diffraction peaks shift to smaller angles with increasing Sm3+ content owing to the different ionic radii between Sm3+ and Lu3+, indicating that Sm3+ ions have been doped into the crystal lattices of the cubic instead of forming the impurity phase. In addition, the Miller indices of Ba3Lu(PO4)3:Sm3+ phosphors were calculated by using equation reported by Tamrakar et al. and given in Table 1 [22]. In order to prove that the actual Sm3+ content corresponds to a nominal composition, the hydrothermal supernatant (30 ml) was measured by inductively coupled high frequency plasma atomic emission spectroscopy (ICP-AES) to detect the residual content of Sm3+ in solution (see Table 2).

 figure: Fig. 3

Fig. 3 XRD patterns of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents. The standard data for Ba3Lu(PO4)3 (JCPDS card no. 43-0212) is shown as a reference.

Download Full Size | PDF

Tables Icon

Table 1. Miller indices value Ba3Lu(PO4)3:Sm3+ phosphor

Tables Icon

Table 2. The doping rate of Sm3+ in the Ba3Lu(PO4)3:x% Sm3+ (x = 1, 3, 5)

3.2 Excitation and emission spectra

To record the luminescence spectra for the prepared phosphor and the possibility for application in w-LEDs, it is necessary to know the excitation wavelengths of Sm3+. For this purpose, we measured the excitation spectrum of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents from 250 to 550 nm monitored at 600 nm. As is shown in Fig. 4(a), it is easy to see there are several excitation peaks of Sm3+, which are located in the wavelength range of 325–500 nm, assigning to the forbidden f-f transition of Sm3+, 6H5/24K17/2 (343 nm), 6H5/24H7/2 (360 nm), 6H5/26P7/2(373 nm), 6H5/24F7/2 (403 nm), 6H5/24G9/2 (438 nm), 6H5/24I11/2 (473 nm) [23, 24]. From the perspective of excitation spectrum, the intensity of transition at 403 nm is higher than any other transitions. This result also indicates that the obtained phosphor can be efficiently excited by NUV chips and will be a potential candidate as light-conversion phosphor for w-LEDs.

 figure: Fig. 4

Fig. 4 (a) Excitation spectra of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents; (b) The diffuse reflectance spectra of Ba3Lu(PO4)3 host, Ba3Lu(PO4)3:0.03Sm3+ and Ba3Lu(PO4)3:0.05Sm3+ phosphors.

Download Full Size | PDF

The reflectance spectra of Ba3Lu(PO4)3 host, Ba3Lu(PO4)3:0.03Sm3+, Ba3Lu(PO4)3: 0.05Sm3+ phosphors are shown in Fig. 4(b). The spectra shows a strong absorption band in the wavelength ranging from 400 to 800 nm, and then shows a remarkable drop from 200 to 350 nm that corresponds to the band transition in the Ba3Lu(PO4)3 host lattice. When Sm3+ ions is introduced into the host lattices, several weak absorption bands in the larger wavelength range 340–500 nm are observed. The absorption range from 200 to 500 nm in the Ba3Lu(PO4)3:Sm3+ phosphor matched well with the photoluminescence excitation spectrum in Fig. 4(a).

The emission spectra of Ba3Lu(PO4)3:xSm3+ (x = 0.01, 0.03, 0.05,0.07 and 0.09) phosphors measured at λex = 403 nm (6H5/24F7/2) excitation wavelength are shown in Fig. 4(a). Three emission peaks were observed at 564, 600 and 647 nm corresponding to 4G5/26H5/2, 4G5/26H7/2, and 4G5/26H9/2, respectively. The emission peak centered at 564 nm (4G5/2→6H5/2) is originated due to purely magnetic dipole moment, second peak at 647 nm (4G5/26H9/2) is due to purely electric dipole moment and the other at 600 nm (4G5/26H7/2) is due to both magnetic and electric dipole moments [25, 26]. Generally, the intensity ratio of electric dipole to magnetic dipole transitions is useful to understand the symmetry of local environment around the trivalent 4f5 ions. In the present phosphor, the magnetic dipole transition is more intense than electric dipole transition suggested that there is no deviation from inversion center and more symmetric in nature [27, 28]. It is observed from Fig. 5(a) that there is no change in the position of emission band for all doping concentrations. However, the emission intensity changes with doping concentration of Sm3+ ions. It is observed the emission intensity of the sample firstly increases with the concentration of Sm3+ increasing. This is because the distance between two adjacent Sm3+ ions is decreased by increasing Sm3+ concentration, and the interaction between Sm3+-Sm3+ ions is enhanced, which causes the nonradiative energy transfer between two Sm3+ ions. The electric multipolar interaction and exchange interaction are involved in the energy transfer [29]. Initially, the emission reaches to a maximum at x = 0.05 (critical concentration), beyond that the emission intensity begins to decrease with increase in concentration due to concentration quenching effect.

 figure: Fig. 5

Fig. 5 (a) Emission spectra of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents; (b) The relationship of lg(x) versus lg(I/x) for Ba3Lu(PO4)3:xSm3+ (x = 0.05, 0.07, 0.09) phosphor.

Download Full Size | PDF

In general, the concentration quenching is mainly caused by non-radiative energy transfer processes (cross-relaxation) among Sm3+ ions, which occurs as a result of an exchange interaction or a multipole–multipole interaction [30]. In order to identify the type of interaction mechanism, it is necessary to obtain the critical distance (Rc) that is the critical separation between the donor (activator) and acceptor (quenching site). According to the theory of Blasse, the critical transfer distance (Rc) of energy transfer can be calculated by the critical concentration of the activator ion [31]:

Rc2[3V4πXcN]1/3
where V is the volume of one unit cell, N is the number of the cationic sites occupied by activators in one unit cell, and Xc is the critical concentration of the activator ion. For Ba3Lu(PO4)3 host, V = 1149.71 Å3, N = 4 and Xc = 0.05. Therefore, by using Eq. (1), the Rc value of Sm3+ can be quickly obtained to be 22.23 Å. According to theory of Van Uitert, if the critical distance between the sensitizer and activator is shorter than 5 Å, the energy transfers type is exchange interaction [32]. The above calculation result suggests that the energy transfer among Sm3+ ions in Ba3Lu(PO4)3:Sm3+ phosphor does not occur in this case. According to Van Uitert, if energy transfer occurs among the same type of activators, the intensity of the multipolar interaction can be determined from the change in the emission intensity from the emitting level that has the multipolar interaction. The emission intensity (I) per activator ion can be expressed as [33]:
Ix=K(1+β(x)Q/3)1
where I is the emission intensity, x is the concentration of the activator ions above the concentration quenching point, β and K are constants for the same conditions, and Q is the function of multipole–multipole interaction for 6 (dipole–dipole), 8 (dipole–quadrupole) or 10 (quadrupole–quadrupole). As shown in Fig. 5(b), to obtain a correct Q the dependence of log(I/x) on log(x) is plotted, and it yields a straight line with a slope of -Q/3. The fitting result for Sm3+ emission centers, which is corresponding to the high Sm3+ concentration Ba3Lu(PO4)3:xSm3+ phosphor compositions, is shown in Fig. 5(b). Based on the fitting results, the slope is −1.66, and the value of Q can be calculated as 4.98, which is approximately equal to 6. This indicates that the dipole–dipole interaction is the major mechanism for concentration quenching in Ba3Lu(PO4)3: Sm3+ phosphor.

3.3 Decay curve and quantum efficiency

For Ba3Lu(PO4)3:xSm3+ phosphors with different concentrations (x = 0.01,0.03, 0.05, 0.07and 0.09), the decay of emission intensity at 600 nm as a function of time is depicted in Fig. 6. The decay curve for 4G5/26H7/2 transition of Sm3+ can be expressed by a double exponential equation as follows [34]:

I(t)=A1exp(t/t1)+A2exp(t/t2)
where I is the luminescence intensity at time t, A1 and A2 are constants, t is the time, and t1 and t2 are the decay time for the exponential components. Then, the average decay lifetime can be calculated as follows:

 figure: Fig. 6

Fig. 6 Decay curves of Ba3Lu(PO4)3:xSm3 + at different concentrations (λex = 370 nm and λem = 600 nm).

Download Full Size | PDF

τ=A1t12+A2t22A1t1+A2t2

According to the Eq. (4), the average lifetimes of Sm3+ ions calculated using the decay profiles are 2.32, 2.19, 1.78, 1.61, and 1.46 ms for Ba3Lu(PO4)3:xSm3+ with x = 0.01, 0.03, 0.05, 0.07 and 0.09, respectively. The lifetime decreases with increased concentration. These results show that the lifetime is short enough for potential applications in displays and lights.

In general, for the practical application of phosphors for LEDs, the quantum efficiency of the phosphor is an important factor to be considered. We have also recorded the quantum yield of Ba3Lu(PO4)3:xSm3+ phosphors. The quantum efficiency can be described by the following equation [35]:

ηQE=λP(λ)dλλ[E(λ)-R(λ)]dλ
where P(λ), E(λ) and R(λ) are the intensities per unit wavelength in the emission, excitation and reflection spectra of the Ba3Lu(PO4)3:xSm3+ phosphor, respectively. Under the excitation of 403 nm, the recorded quantum yield values of Ba3Lu(PO4)3:xSm3+ are 0.14, 0.31, 0.48, 0.24, and 0.19 for x = 0.01, 0.03, 0.05, 0.07, and 0.09 respectively, which are consistent with the variation of the emission intensities.

3.4 CIE coordinates and thermal properties

Thermal stability is an important technological parameter for phosphors used in solid-state lighting, since it exerts significant influence on the light output as well as color-rendering index. The temperature-dependent emission spectra of the Ba3Lu(PO4)3:0.05Sm3+ phosphor measured at 25 °C–300 °C under 403 nm excitation were shown in Fig. 7(a). One can see that the intensity of emission spectrum gradually decreases with the increase of temperature due to the thermal quenching effect. As shown in Fig. 7(b), when the measurement temperature is up to 250 °C, the emission intensities of 4F9/26H13/2 decrease to 84.6% of the initial emission intensities measured at 25 °C, which is higher than that of Y3Al5O12:Ce3+ commercial phosphor (58%) [36].

 figure: Fig. 7

Fig. 7 (a) Temperature dependence of the PL intensity of Ba3Lu(PO4)3:0.05Sm3+ phosphor; (b) The relative emission intensity of Ba3Lu(PO4)3:0.05Sm3+ and YAG:Ce as a function of temperature; (c) Activation energy for thermal quenching of Ba3Lu(PO4)3:0.05Sm3+ phosphor.

Download Full Size | PDF

To further investigate the thermal quenching phenomenon and obtain the activation energy for thermal quenching, the activation energy from the thermal quenching can be calculated using the Arrhenius equation [37]:

I(T)=Io1+cexp((ΔE/KT))
where I0 is the initial intensity, I(T) is the intensity at a given temperature T, ΔE is the activation energy for thermal quenching, c is a constant for a certain host, and K is the Boltzmann constant (8.629 × 10−5 eV). Therefore, we used the Eq. (6) to fit the temperature dependent luminescent intensities shown in Fig. 7(b). During the fitting process, the activation energy (ΔE) was deduced to be 0.142 eV. This result indicates that the as-prepared phosphors possess good thermal stability properties.

Color coordinate is also one of the important factors for evaluating phosphors’ performance. The Commission Internationale de L’Eclairage (CIE) chromaticity diagram of the prepared Ba3Lu(PO4)3:0.05Sm3+ phosphor is shown in Fig. 8(a). The color coordinate for Ba3Lu(PO4)3:0.05Sm3+ under 403 nm excitation is shown as (0.602, 0.389), which means that the emitted light is in the region of orange reddish light. In order to investigate the possibility of the as-synthesized phosphor in application for NUV LED, a prototype red LED was fabricated by pre-coating Ba3Lu(PO4)3:0.05Sm3+ phosphor onto a 404 nm emitting InGaN chip. The emission spectra of the original 404 nm emitting InGaN chip and prototype red LED under 20 mA forward bias current are shown in Fig. 8(b). The band at 404 nm results from emission of InGaN chip. When coated with Ba3Lu(PO4)3:0.05Sm3+ phosphor, the near UV light was absorbed by the phosphor, and simultaneously down-converted into an intensive red light. The results demonstrated that Ba3Lu(PO4)3:0.05Sm3+ phosphor could efficiently absorb the NUV light, thereby making it a good candidate as red component of NUV LED application.

 figure: Fig. 8

Fig. 8 (a) The emission spectra of red pc-LED (Inset: photo of red LED under a bias current of 20 mA); (b) Chromaticity coordinates of Ba3Lu(PO4)3:0.05Sm3+ phosphor in the CIE 1931 chromaticity diagram.

Download Full Size | PDF

4. Conclusions

In summary, the novel phosphor of Ba3Lu(PO4)3 doped with rare earth ions of Sm3+ were synthesized by a solid-state reaction. The PLE spectra shows that the phosphor can be efficiently excited by ultraviolet light and emit a satisfactory orange-red performance, nicely, fitting in well with the widely used NUV LED chip. The thermal stability of Ba3Lu(PO4)3:0.05Sm3+ sample is superior to the commercial YAG:Ce3+ phosphor. All the obtained results indicate that these phosphors might be a promising reddish-orange emitting phosphor used in w-LEDs.

Funding

National Natural Science Foundation of China (NSFC) (21576002).

References and links

1. S. J. Ho and Y. J. Duk, “Yellow-emitting Sr3SiO5:Ce3+, Li+ phosphor for white–light emitting diodes and yellow-light-emitting diodes,” Appl. Phys. Lett. 90(4), 041906 (2007). [CrossRef]  

2. P. Pust, V. Weiler, C. Hecht, A. Tücks, A. S. Wochnik, A. K. Henß, D. Wiechert, C. Scheu, P. J. Schmidt, and W. Schnick, “Narrow-band red-emitting Sr[LiAl3N4]:Eu2+ as a next-generation LED-phosphor material,” Nat. Mater. 13(9), 891–896 (2014). [CrossRef]   [PubMed]  

3. M. Y. Chen, Z. G. Xia, M. S. Molokeev, and Q. L. Liu, “Insights into Ba4Si6O16 structure and photoluminescence tuning of Ba4Si6O16:Ce3+, Eu2+ phosphors,” J. Mater. Chem. C 3(48), 12477–12483 (2015). [CrossRef]  

4. B. Ramesh, G. Devarajulu, B. D. P. Raju, G. B. Kumar, G. R. Dillip, A. N. Banerjee, and S. W. Joo, “Determination of strain, site occupancy, photoluminescent, and thermoluminescent-trapping parameters of Sm3+-doped NaSrB5O9 microstructures,” Ceram. Int. 42(1), 1234–1245 (2016). [CrossRef]  

5. F. A. Ponce and D. P. Bour, “Nitride-based semiconductors for blue and green light -emitting devices,” Nature 386(6623), 351–359 (1997). [CrossRef]  

6. C. Liang, H. You, Y. Fu, X. Teng, K. Liu, and J. He, “A novel tunable blue-green-emitting CaGdGaAl2O7:Ce(3+),Tb(3+) phosphor via energy transfer for UV-excited white LEDs,” Dalton Trans. 44(17), 8100–8106 (2015). [CrossRef]   [PubMed]  

7. X. G. Zhang, J. L. Zhang, and M. L. Gong, “Synthesis and luminescent properties of UV-excited thermal stable red-emitting phosphor Ba3Lu(PO4)3: Eu3+ for NUV LED,” Opt. Mater. 36(4), 850–853 (2014). [CrossRef]  

8. R. Chen, Y. H. Hu, Y. H. Jin, L. Chen, and X. J. Wang, “Tunable emission and efficient energy-transfer properties of Ce3+ and Mn2+ co-doped Ba3Gd(PO4)3 phosphors,” Appl. Phys. A. 117(2), 823–829 (2014). [CrossRef]  

9. J. Barbier, “Structural refinements of eulytite-type Ca3Bi (PO4)3 and Ba3La(PO4)3,” J. Solid State Chem. 101(2), 249–256 (1992). [CrossRef]  

10. Q. M. Di, Q. G. Xu, L. Han, and J. Y. Sun, “Synthesis and luminescence properties of Eu3+-doped Ba3Gd(PO4)3 phosphors for light-emitting diodes,” Opt. Eng. 54(3), 035104 (2015). [CrossRef]  

11. H. Zhu, C. C. Lin, W. Luo, S. Shu, Z. Liu, Y. Liu, J. Kong, E. Ma, Y. Cao, R. S. Liu, and X. Chen, “Highly efficient non-rare-earth red emitting phosphor for warm white light-emitting diodes,” Nat. Commun. 5, 4312 (2014). [CrossRef]   [PubMed]  

12. C. J. Duan, A. C. A. Delsing, and H. T. Hintzen, “Photoluminescence properties of novel red-emitting Mn2+-activated MZnOS (M= Ca, Ba) phosphors,” Chem. Mater. 21(6), 1010–1016 (2009). [CrossRef]  

13. B. Chen, Q. Zhou, J. Li, F. Zhang, R. Liu, H. Zhong, and B. Zou, “Red emissive CuInS2-based nanocrystals: a potential phosphor for warm white light-emitting diodes,” Opt. Express 21(8), 10105–10110 (2013). [CrossRef]   [PubMed]  

14. A. K. Vishwakarma and M. Jayasimhadri, “Pure orange color emitting Sm3+ doped BaNb2O6 phosphor for solid - state lighting applications,” J. Lumin. 176, 112–117 (2016). [CrossRef]  

15. B. F. Lei, Y. L. Liu, G. B. Tang, Z. R. Ye, and C. S. Shi, “Spectra and long-lasting properties of Sm3+-doped yttrium oxysulfide phosphor,” Mater. Chem. Phys. 87(1), 227–232 (2004). [CrossRef]  

16. F. Cheng, Z. Xia, M. S. Molokeev, and X. Jing, “Effects of composition modulation on the luminescence properties of Eu3+ doped Li1-xAgxLu(MoO4)2 solid-solution phosphors,” Dalton Trans. 44(41), 18078–18089 (2015). [CrossRef]   [PubMed]  

17. N. Guo, Y. Huang, Y. Jia, W. Lv, Q. Zhao, W. Lü, Z. Xia, and H. You, “A novel orange-yellow-emitting Ba3Lu(PO4)3:Eu2+,Mn2+ phosphor with energy transfer for UV-excited white LEDs,” Dalton Trans. 42(4), 941–947 (2013). [CrossRef]   [PubMed]  

18. C. Jin, H. X. Ma, Y. F. Liu, Q. B. Liu, G. Y. Dong, and Q. M. Yu, “Tunable luminescence properties and energy transfer in Ba3Lu(PO4)3:Ce3+,Tb3+ phosphors,” J. Alloys Compd. 613(1), 275–279 (2014). [CrossRef]  

19. H. B. Liang, Y. Tao, J. H. Xu, H. He, H. Wu, W. X. Chen, S. B. Wang, and Q. Su, “Photoluminescence of Ce3+, Pr3+ and Tb3+ activated Sr3Ln(PO4)3 under VUV-UV excitation,” J. Solid State Chem. 177(3), 901–908 (2004). [CrossRef]  

20. N. Guo, Y. H. Zheng, Y. C. Jia, H. Qiao, and H. P. You, “Warm-white-emitting from Eu2+/Mn2+-codoped Sr3Lu(PO4)3 phosphor with tunable color tone and correlated color temperature,” J. Phys. Chem. C 116(1), 1329–1334 (2012). [CrossRef]  

21. M. M. Yawalkar, G. D. Zade, K. V. Dabre, and S. J. Dhoble, “Luminescence study of Eu3+ doped Li6 Y(BO3 )3 phosphor for solid-state lighting,” Luminescence 31(4), 1037–1042 (2016). [CrossRef]   [PubMed]  

22. R. K. Tamrakar, “Thermoluminescence studies of copper-doped cadmium sulphide nanoparticles with trap depth parameters,” Res. Chem. Intermed. 39(9), 4239–4245 (2013). [CrossRef]  

23. Z. W. Zhang, Y. S. Peng, X. H. Shen, J. P. Zhang, S. T. Song, and Q. Lian, “Enhanced novel orange red emission in LiSr4−x(BO3)3:xSm3+ by K+,” J. Mater. Sci. 49(6), 2534–2541 (2014). [CrossRef]  

24. Z. P. Yang, Y. Han, Y. C. Song, Y. H. Zhao, and P. F. Liu, “Synthesis and luminescence properties of a novel red Sr3Bi (PO4) 3: Sm^ 3+ phosphor,” J. Rare Earths 30(12), 1199–1202 (2012). [CrossRef]  

25. G. S. R. Raju and S. Buddhudu, “Emission analysis of Sm3+ and Dy3+:MgLaLiSi2O7 powder phosphors,” Spectrochim. Acta A Mol. Biomol. Spectrosc. 70(3), 601–605 (2008). [CrossRef]   [PubMed]  

26. V. Singh, S. Watanabe, T. K. G. Rao, J. F. D. Chubaci, and H. Y. Kwak, “Luminescence and defect centres in MgSrAl10O17:Sm3+ phosphor,” J. Non-Cryst. Solids 356(3), 1185–1190 (2010). [CrossRef]  

27. Z. G. Xia and D. M. Chen, “Synthesis and Luminescence Properties of BaMoO4:Sm3+ phoshpors,” J. Am. Ceram. Soc. 93(5), 1397–1401 (2010).

28. Z. F. Yang, D. H. Xu, J. Y. Sun, Y. M. Sun, and H. Y. Du, “Characterization and luminescence properties of Sr3Gd(PO4)3:Sm3+ orange-red phosphor,” Opt. Eng. 54(10), 105102 (2015). [CrossRef]  

29. Y. Q. Zhai, W. Zhang, Y. J. Yin, Y. Han, X. Zhao, H. H. Ding, and N. Li, “Morphology tunable synthesis and luminescence property of NaGd(MoO4)2:Sm3+ microcrystals,” Ceram. Int. 43(1), 841–846 (2017). [CrossRef]  

30. M. F. Joubert, A. Remillieux, B. Jacquier, J. Mugnier, B. Boulard, O. Perrot, and C. Jacoboni, “Infrared to visible conversion in rare-earth-doped planar waveguides,” J. Non-Cryst. Solids 184(1), 341–345 (1995). [CrossRef]  

31. G. Blasse, “Energy transfer between inequivalent Eu2+ ions,” J. Solid State Chem. 62(2), 207–211 (1986). [CrossRef]  

32. L. G. Van Uitert, “Characterization of energy transfer interactions between rare earth ions,” J. Electrochem. Soc. 114(10), 1048–1053 (1967). [CrossRef]  

33. Z. F. Yang, D. H. Xu, J. N. Du, X. D. Gao, and J. Y. Sun, “Tunable luminescence and energy transfer of Eu2+/Mn2+ co-doped Sr3NaY(PO4)3F phosphor for white LEDs,” RSC Advances 6(90), 87493–87501 (2016). [CrossRef]  

34. J. H. Hao, S. A. Studenikin, and M. Cocivera, “Transient photoconductivity properties of tungsten oxide thin films prepared by spray pyrolysis,” J. Appl. Phys. 90(10), 5064–5069 (2001). [CrossRef]  

35. H. Zhou, Q. Wang, M. Jiang, X. Jiang, and Y. Jin, “A novel green-emitting phosphor Ba2Gd2Si4O13:Eu(2+) for near UV-pumped light-emitting diodes,” Dalton Trans. 44(31), 13962–13968 (2015). [CrossRef]   [PubMed]  

36. D. H. Xu, Z. F. Yang, J. Y. Sun, X. D. Gao, and J. N. Du, “Synthesis and luminescence properties of double-perovskite white emitting phosphor Ca3WO6:Dy3+,” J. Mater. Sci. Mater. Electron. 27(8037), 8370–8377 (2016). [CrossRef]  

37. W. R. Liu, C. H. Huang, C. W. Yeh, J. C. Tsai, Y. C. Chiu, Y. T. Yeh, and R. S. Liu, “A study on the luminescence and energy transfer of single-phase and color-tunable KCaY(PO4)2:Eu2+,Mn2+ phosphor for application in white-light LEDs,” Inorg. Chem. 51(18), 9636–9641 (2012). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1
Fig. 1 (a) Structural view of Ba3Lu(PO4)3; (b) The coordination environment of Ba2+/Lu3+ site.
Fig. 2
Fig. 2 SEM images of Ba3Lu(PO4)3:0.05Sm3+ sample under different magnification (a): 2.2 k × ; (b) 3.5 k × .
Fig. 3
Fig. 3 XRD patterns of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents. The standard data for Ba3Lu(PO4)3 (JCPDS card no. 43-0212) is shown as a reference.
Fig. 4
Fig. 4 (a) Excitation spectra of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents; (b) The diffuse reflectance spectra of Ba3Lu(PO4)3 host, Ba3Lu(PO4)3:0.03Sm3+ and Ba3Lu(PO4)3:0.05Sm3+ phosphors.
Fig. 5
Fig. 5 (a) Emission spectra of Ba3Lu(PO4)3:Sm3+ with different Sm3+ contents; (b) The relationship of lg(x) versus lg(I/x) for Ba3Lu(PO4)3:xSm3+ (x = 0.05, 0.07, 0.09) phosphor.
Fig. 6
Fig. 6 Decay curves of Ba3Lu(PO4)3:xSm3 + at different concentrations (λex = 370 nm and λem = 600 nm).
Fig. 7
Fig. 7 (a) Temperature dependence of the PL intensity of Ba3Lu(PO4)3:0.05Sm3+ phosphor; (b) The relative emission intensity of Ba3Lu(PO4)3:0.05Sm3+ and YAG:Ce as a function of temperature; (c) Activation energy for thermal quenching of Ba3Lu(PO4)3:0.05Sm3+ phosphor.
Fig. 8
Fig. 8 (a) The emission spectra of red pc-LED (Inset: photo of red LED under a bias current of 20 mA); (b) Chromaticity coordinates of Ba3Lu(PO4)3:0.05Sm3+ phosphor in the CIE 1931 chromaticity diagram.

Tables (2)

Tables Icon

Table 1 Miller indices value Ba3Lu(PO4)3:Sm3+ phosphor

Tables Icon

Table 2 The doping rate of Sm3+ in the Ba3Lu(PO4)3:x% Sm3+ (x = 1, 3, 5)

Equations (6)

Equations on this page are rendered with MathJax. Learn more.

R c 2 [ 3V 4π X c N ] 1/3
I x =K ( 1+β ( x ) Q/3 ) 1
I( t )= A 1 exp( t/ t 1 )+ A 2 exp( t/ t 2 )
τ= A 1 t 1 2 + A 2 t 2 2 A 1 t 1 + A 2 t 2
η QE = λP( λ )dλ λ[E(λ)-R(λ)]dλ
I( T )= I o 1+cexp( ( ΔE/KT ) )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.